Generation of Nanoparticles with Adjustable Size and Controlled

Jul 12, 2012 - Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Científicas (CSIC), C/Sor Juana Inés de la Cruz,. 3, ...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/Langmuir

Generation of Nanoparticles with Adjustable Size and Controlled Stoichiometry: Recent Advances L. Martínez, M. Díaz, E. Román, M. Ruano, D. Llamosa P., and Y. Huttel* Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Científicas (CSIC), C/Sor Juana Inés de la Cruz, 3, 28049 Madrid, Spain ABSTRACT: We present a bottom-up fabrication route based on the sputtering gas aggregation source that allows the generation of nanoparticles with controllable and tunable chemical composition while keeping the control of the cluster size. We demonstrate that the chemical composition of the particles can be monitored by the individual adjustment of the working parameters of the magnetrons inserted in a gas aggregation zone. Such control of the parameters leads to a fine control of the ion density of each target material and hence to the control of the chemical composition of the nanoparticles. In particular, we show through X-ray photoemission, atomic force microscopy, and high-resolution transmission electron microscopy that it is possible to generate bimetallic (AgAu) and trimetallic (AgAuPd) alloy nanoparticles with welldefined and tunable stoichiometries from three targets of pure Ag, Au, and Pd. The proposed route for the generation of nanoparticles opens new possibilities for the fabrication of nanoparticles using a physical method that, for some applications, could be complementary to the chemical methods. commercially available in 2001,17 and it is relatively easy to use. Other reasons for the popularity of this kind of NP source probably relays on the fact that the sputtering process is wellknown in the industry and that it generates an important proportion of ionized NPs that can be mass/charge selected by the attachment of a quadrupole to the ion cluster source (ICS). In fact, since the initial developments of the group of Haberland,10,11 Binns and collaborators demonstrated that the use of a quadrupole to charge select the NPs is very effective in the reduction of the dispersion of the size distribution of the NPs.18,19 The standard ICS is composed by a single magnetron (usually 2 in. magnetron) that generates ions of a given material that aggregates in the so-called aggregation zone. The size of the clusters (typically 1−20 nm) is controlled by the power applied to the magnetron, the argon flux (for the sputtering process), the position of the magnetron in the aggregation zone, and the flux of an extra gas that is usually helium. The flux of clusters or NPs that exit the ICS through an aperture is controlled by the time. The ICS is usually connected to a vacuum or ultrahigh vacuum chamber where a substrate is placed and collects the NPs. In such a design, the chemical composition of the particles depends on the chemical composition of the target that is placed in the magnetron. It appears that, in most cases, the chemical composition of the clusters is close or identical to that of the target material.16 Several modifications to the ICS have been reported in the literature. Most of them have been oriented to the generation

I. INTRODUCTION Among the physical methods that are used for the fabrication of nanoparticles (NPs), the gas aggregation source is attracting special interest in the scientific and industrial sectors due to its inherent specificities. Many characteristics of the gas aggregation sources make them particularly well suited for given applications. In particular, it should be mentioned the ability to work in vacuum or ultrahigh vacuum (UHV) conditions that not only guarantees the chemical purity of the NPs but also is mandatory in some specific applications.1−3 Also, the size of the NPs can be monitored and the composition can be altered in situ by the introduction of gases such as oxygen and nitrogen in order to form oxidized and nitrided NPs with a tunable stoichiometry. Finally, it should be noted that, with this technique, the substrate or surface conditions where the particles are deposited are independent of the fabrication of the NPs. This is due to the fact that they are formed before their landing on the substrate that can be of any material, kept at any temperature, etc. Since the initial developments of the cluster sources,4−6 different designs have been reported with enhanced capabilities.7−11 All the gas aggregation sources are based on the formation of a vapor made of elements that aggregate in order to form NPs. They differ on the way the vapor of the elements is produced: arc discharge, pulsed microplasma, laser vaporization, Joule heating, and sputtering. The present paper is not intended to be a review paper on the gas aggregation sources, and we refer the reader to already published reviews on the subject.12−16 We will now focus on the gas aggregation source that is based on sputtering. Such type of gas aggregation source is one of the most popular both in the laboratories and in the industry. This is probably related to the fact that it became © 2012 American Chemical Society

Received: May 31, 2012 Revised: July 11, 2012 Published: July 12, 2012 11241

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

Figure 1. Scheme of the multiple ion cluster source that is operated with three individual 1 in. diameter magnetrons.

of high cluster fluxes,20−22 postannealing treatments for the generation of magnetic particles with appropriate crystallographic structure, and high magnetic anisotropy.23−25 Other modifications of the ICS have been developed in order to control the oxidation of clusters,26 to focus the cluster beam by means of aerodynamic lenses,27 to control the landing energy of the clusters on the substrates,28 and to allow the room temperature operation of the ICS,29 among others. Surprisingly, only few works have been reported in the literature on modifications of the ICS in order to add an in-situ control and adjustment of the chemical composition of the generated NPs.30−32 While Pellarin et al.30,31 have used a double ablation laser to generate (C60)nSi+m cationic clusters by quenching the vapors from two independent C60 and silicon targets, Sumiyama et al.32 used two sputtering sources to produce Co/CoO and Co/Si core−shell clusters. In both designs, the positions of the targets are fixed and hence cannot be used as a parameter to modify the NPs size like in a more standard ICS.21,33 As far as we know, there is no report on the use of two or more independent target materials to produce alloyed particles in an ICS. In most ICS, the chemical composition of the fabricated NPs depends on the chemical composition of the single target material. The modification of the chemical composition of the NPs implies the substitution of the single target by another one with a different chemical composition. The modification of the ICS that we present in this work allows the generation of NPs with controlled and tunable chemical composition without losing the size control and avoiding the change of the target placed into the ICS. This last point would be of great interest for fundamental studies but also for application purposes in catalysis and information technology, for example.34,35 We have therefore dedicated efforts in developing an ICS that allows the fine-tuning of the chemical composition of the NPs without the need to change the target material for each desired chemical composition.36 In the present paper we show that the insertion of three independent magnetrons into the aggregation zone of an ICS allows the generation of NPs with adjustable chemical composition and controlled stoichiometry without the need to change the target materials. It is shown by atomic force microscopy (AFM) measurements that the new design presented here allows the fabrication of size controlled NPs made of Ag, Au, and Pd. X-ray photoemission (XPS) measurements reveal that the NPs are made of bimetallic AuAg and trimetallic AuAgPd alloys. The chemical composition is controlled by adjusting the working conditions of the magnetrons. Additionally, high-resolution transmission electron

microscopy (HRTEM) performed on individual NPs clearly reveals the alloyed nature of the generated nanoparticles.

II. EXPERIMENTAL DETAILS The modified ICS that we tentatively named multiple ion cluster source (from now on MICS) was attached to an UHV chamber with a base pressure in the middle 10−9 mbar. The clusters generated with the MICS were deposited on flat silicon wafers (10 × 10 mm2) that were transferred into the chamber through a fast entry load lock. The working distance (distance between the exit slit of the MICS and the substrate placed into the manipulator of the chamber) was ∼20 cm. AFM measurements have been performed using the Cervantes AFM System equipped with the “Dulcinea” electronics from Nanotec Electronica S.L.37 in the dynamic mode using commercial silicon AFM tips with a typical radius less than 7 nm. Images were also recorded using modified tips with enhanced lateral resolution. For that purpose commercial tips were covered with 2−3 nm diameter NPs that induce an aspect-ratio and lateral-resolution enhancement in AFM.38,39 The WSxM software from Nanotec40 and Gwyddion software have been used for the analysis of the images. XPS measurements were performed in a separate UHV chamber with a base pressure of 2 × 10−10 mbar, using a Phoibos 100 ESCA/ Auger spectrometer with a Mg Kα anode (1253.6 eV). The samples were soft Ar+ sputtered in order to partially remove the air contamination resulting from the sample transfer from the MICS to the XPS experimental setup. XPS spectra were recorded before and after the soft Ar+ sputtering to check possible sputtering induced intermixing. The absence of such effect was carefully checked. The Au 4f, Ag 3d, and Pd 3d core level narrow spectra were recorded using a pass energy of 15 eV and an energy step of 50 meV. For the data analysis, the contributions of the Mg Kα satellite lines were subtracted, and the spectra were subjected to a Shirley background subtraction formalism. The binding energy (BE) scale was calibrated with respect to the C 1s core level peak at 285 eV. HRTEM measurements were performed with a FEI-TITAN X-FEG transmission electron microscope used in STEM mode and operated at 300 kV. The images were acquired using a high angle annular dark field (HAADF) detector with the desired camera length easily satisfying the Z2 contrast condition. In addition, the microscope is equipped with monochromator, Gatan Energy Filter Tridiem 866 ERS, and a spherical aberration corrector (CEOS) for the electron probe, allowing an effective 0.08 nm spatial resolution. The column is also fitted with energy dispersive X-ray spectroscopy (EDX).

III. RESULTS AND DISCUSSION A. Multiple Ion Cluster Source Description. Figure 1 displays a scheme of the MICS. The three magnetrons are mounted on a 6 in. flange and are inserted into the aggregation zone of a NC200U-B model ion cluster source from Oxford Applied Research Ltd.17 The 1 in. magnetrons have been 11242

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

fabricated following the original design of Prof. Jose M. Colino Garciá from the Facultad de Ciencias del Medio Ambiente, Toledo, Spain.41 As can be observed, each magnetron possess a translation motion that allows its individual positioning inside the aggregation zone. With such a design, the main translation motion of the NC200U-B model ion cluster source is preserved and can be used to vary the size of the clusters like in a standard ICS. In the scheme presented in Figure 1 the magnetrons are placed at the same distance from the exit slit of the MICS. We will show that, within this configuration, the formation of alloy NPs is achievable. This has been illustrated by the sputtering of ions of given materials (red and blue) that further mix and aggregate forming alloy NPs (green). Each magnetron is provided with its own argon flux, cooling pipes, and electrical connection. An additional gas entry is also available for the injection of another gas in the aggregation zone (oxygen in order to oxidize the NPs or helium in order to reduce the average size of the NPs for example). Note that, as in the case of the standard ICS, the different parameters that allow a control on the size of the NPs (power applied to the magnetrons, argon flux, positions of the magnetrons in the aggregation zone, and the helium flux) are kept in the design of the MICS but individually for each magnetron, which makes them completely independent of each other. As the control of the size of the NPs is basically the same as in standard ICS, we do not present a detailed study of that aspect and rather focus on the control of the chemical composition of the NPs that is the novelty of this equipment. Under this design, the formation of alloy NPs is achievable by controlling the density of ions of each magnetron that is monitored through the applied power and the argon flux. In Figure 2a we show a photograph of the MICS attached to the NC200U-B model ion cluster source. The main translator of the latter and the 3 individual translators of the MICS are indicated. Figure 2b presents a photograph of the three individual 1 in. diameter magnetrons before their introduction into the aggregation zone. B. Production of the Alloy Nanoparticles. All the deposits have been performed with the magnetrons positioned at a fixed position with respect to the exit slits of the aggregation zone. For the present study we have used three magnetrons with silver (99.99%), gold (99.999%), and palladium (99.99%) targets. At the indicated working distance (≈20 cm) the spot diameter of the area where the NPs landed was ≈40 mm. A series of samples have been produced, keeping the same fluxes of argon in the magnetron and without additional gas flux. The only parameter that has been adjusted is the power applied to each magnetron. The deposition times were fixed in order to produce multilayers of NPs suitable for XPS characterization and also to produce submonolayer assembly of NPs for appropriate AFM and HRTEM measurements. The chemical composition derived from the XPS measurements together with the applied powers and argon fluxes are given in Tables 1 and 2. Two series of alloyed NPs were performed. The first one consisted of bimetallic AuAg alloyed NPs with five different stoichiometries, from pure Au to pure Ag, passing through three intermediate chemical compositions. This series clearly demonstrates the ability to precisely control the stoichiometry of the fabricated NPs. In the second series, AuAgPd NPs are fabricated in order to demonstrate that it is possible to fabricate trimetallic NPs without losing control over the stoichiometry. This ability to control the chemical composition is particularly interesting in

Figure 2. (a) Photograph of the multiple ion cluster source connected to the NC200U-B model ion cluster source. (b) Photograph of the three individual 1 in. diameter magnetrons.

Table 1. Powers and Argon Fluxes Applied to the Au and Ag Magnetronsa samples

PAg (W)

ArAg (sccm)

PAu (W)

ArAu (sccm)

Ag100Au0 Ag74Au26 Ag55Au45 Ag35Au65 Ag0Au100

14.3 8.0 7.7 5.0 0

40 40 40 40 40

0 6.1 9.3 9.3 9.3

40 40 40 40 40

a

The indicated chemical composition is the one extracted from the XPS data (cf. XPS Characterization section).

Table 2. Powers Applied to the Au, Ag, and Pd Magnetronsa samples

PAg (W)

PAu (W)

PPd (W)

Ag0Au0Pd100 Ag32Au42Pd26 Ag27Au17Pd56

0 7.6 4.9

0 9.3 5.8

12 7.3 9.8

a

The indicated chemical composition is the one extracted from the XPS data (cf. XPS Characterization section). All the Ar fluxes were fixed to 30 sccm.

order to generate NPs with the desired properties, like catalytic42 or magnetic43,44 properties for specific applications. In our case, as the NPs are fabricated in a controlled 11243

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

atmosphere (UHV conditions), they present a high purity with very limited amount of contaminants. A representative scanning electron microscope (SEM) image of a multilayer of nanoparticles is displayed in Figure 3. Both

Figure 3. Representative SEM image of a multilayer of Ag50Au50 nanoparticles deposited on a silicon wafer.

the flat silicon wafer substrate and the multilayer of nanoparticles can be distinguished in the SEM image. The observed edge corresponds to the shadow of the sample holder where the silicon wafer was fixed. Although the multilayer of nanoparticles appears quite compact and flat, some roughness can be clearly observed in Figure 3. Typical deposition times for such multilayers of ∼350 nm were 4 min which clearly indicates the high efficiency of the process. Moreover, individual nanoparticles can be distinguished in the silicon wafer region. Higher magnification topography images have been recorded by means of AFM and are presented in the following section. C. AFM Characterization of the Deposits. Figure 4 shows the AFM images recorded in all the deposits. While Figures 4a−e correspond to AuAg nanoparticles, Figures 4f−h correspond to AuAgPd nanoparticles. It can be clearly observed that all deposits are made of NPs forming multilayered systems. The NP size was kept in the same range for all samples in order to avoid size effects, if any, in XPS experiments (cf. next section). Nevertheless, some dispersion of diameters can be observed in Figure 4. Such dispersion is related to changes in the power applied to the magnetrons for the fabrication of each sample. Although not performed in this study, the dispersion in diameter can be corrected by the use of a quadrupole18,19 or by modifying the aggregation length and/or injecting a helium flux in the aggregation zone. Figure 5 corresponds to a submonolayer assembly of NPs of chemical composition Ag55Au45. The submonolayer was achieved by reducing the deposition time to a few seconds (typically 5 s). As can be observed in Figure 5a, the number of deposited NPs is still high (typically more than 150 NPs/μm2) which illustrates once more the high efficiency of the fabrication process. Since the substrates were not cooled down or heated, it is very improbable that the formation of the NPs would result from a Stranski−Krastanov-like growth. The AFM images that illustrate the deposition of individual NPs clearly show that, like in standard ICS, the NPs are generated into the MICS and subsequently deposited on the desired substrate. Figure 5b

Figure 4. AFM images acquired on the NPs deposited on a silicon surface for different powers applied to the magnetrons (cf. Tables 1 and 2). The chemical compositions of each of the deposits extracted from the XPS measurements (cf. next section) are also indicated. Parts a−e correspond to multilayers of AuAg NPs, and parts f−h correspond to AuAgPd NPs.

represents the height distribution of the NPs. This distribution has been extracted from the analysis of several AFM images and includes a total number of NPs close to 700. It has been fitted using a Galton or log-normal distribution.45,46 The resulting mean height and standard deviation were found to be 6.3 and 1.1 nm, respectively. The standard deviation is similar to the one obtained when using a single magnetron (e.g., 6.3 ± 0.7 nm for pure Ag nanoparticles and 6.1 ± 1.2 nm for pure Au nanoparticles). Moreover, it is in the same range as previous analysis carried out on NPs fabricated with a standard ICS without a quadrupole filter at the end of the aggregation zone.46 11244

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

Figure 6. (a) Ag 3d and (b) Au 4f core level XPS spectra of the NPs grown with different stoichiometry of Ag and Au.

presence of a dense assembly of Au NPs.47 It can be also observed that the binding energy of the Au 4f7/2 core level peak decreases with decreasing content of Au in the NPs. The evolution of the chemical shifts of the Ag 3d5/2 and Au 4f7/2 core level peaks with respect to the bulk metallic Ag and Au core level peaks have been represented as a function of the stoichiometry of the NPs in Figure 7. The presented chemical shifts compare reasonably well with the previous studies of Harikumar et al.,47 hence supporting the presence of NPs of AgAu alloy. It is important to notice that no Au was detected in the pure Ag sample (Ag100Au0) and no Ag was detected in the pure Au sample (Ag0Au100). This implies that the magnetrons are not contaminating each other. This is an important issue since the contamination from neighbor targets can happen when plasmas from different magnetrons interact which is not the case here. As it was previously mentioned, each magnetron has its own power supply and gas entry and works independently from each other. When fabricating pure metal NPs from one single magnetron, the power supplies of the other magnetrons are turned off, avoiding any possible contamination. However, when performing alloys, more than one magnetron is turned on. The individual vapors mix in the aggregation zone in a proportion that is given by the power applied to each magnetron and in such way that is possible to control the stoichiometry of the fabricated NPs. 2. AuAgPd Nanoparticles. Similar XPS characterization has been performed on the AuAgPd NPs. The Ag 3d5/2, Pd 3d5/2, and Au 4f7/2 core level XPS spectra are displayed in Figure 8. The BE positions corresponding to pure Ag, Pd, and Au NPs have been included in the figures as dashed lines for

Figure 5. (a) AFM image of a submonolayer assembly of Ag55Au45 alloyed NPs. (b) Height distribution extracted from several AFM images.

D. XPS Characterization of the Deposits. 1. AuAg Nanoparticles. Further characterization of the deposits has been done with XPS. The Ag 3d and Au 4f core level spectra are displayed in Figure 6. The chemical composition of the NPs has been determined analyzing the integrated intensities of Au 4f and Ag 3d core level peaks, following the procedure explained in the Experimental Details section and taking into account the appropriate relative sensitivity factors. The measured stoichiometry of the samples is reported in Figure 6. Offsets have been applied to the spectra in order to facilitate the comparison between spectra and the intensity is given in counts per second. The spectra clearly show changes in the intensities in agreement with the modification of the chemical composition or stoichiometry of the NPs. The BE of the Ag 3d5/2 core level peak (Figure 6a) that corresponds to pure Ag NPs (368.7 eV) is located at higher BE than for bulk metallic Ag (368.2 eV). This energy shift is coherent with the presence of NPs. In fact, it is well-known that the binding energy position of the Ag 3d5/2 peak in AgAu and Ag NP depends on the NP size, coverage, and stoichiometry.47−50 Also observable is the slight progressive shift toward lower BE with the increasing Au content in agreement with previous studies on AgAu NPs.47−49 The corresponding evolution of the Au 4f core level can be followed in Figure 6b. The measured BE of pure Au NPs in the Au 4f7/2 core level peak (84.3 eV) is located at higher BE than for bulk metallic Au (84.0 eV), which is in agreement with the 11245

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

Figure 7. Evolutions of the chemical shifts of the Ag 3d5/2 and Au 4f7/2 core level peaks with respect to the bulk metallic Ag and Au.

comparison purposes. The aim of this work is not to analyze in depth the evolution of the chemical shifts of the core level peaks of the different elements, but to demonstrate the ability to produce alloyed nanoparticles with tunable stoichiometry by using the MICS. The fine analysis of the chemical shifts goes beyond the scope of the present paper. Furthermore, there are only few studies on AuAgPd alloyed NPs reported in the literature,51,52 and none of them present XPS results that could be compared with our results. The stoichiometry extracted from the XPS data is the one given in Table 2 and also reported in Figure 8. The chemical composition has been extracted from the integrated intensities of the XPS core level spectra. Special care has been paid to the overlapping of the Pd 4s peak with the Au 4f peaks and Pd 3d with Au 4d5/2 during the quantification analysis. From Figure 8, it can be seen that no Au or Ag are detected in the sample made of pure Pd NPs, indicating once again that the magnetrons are not contaminating each other. The BE of the Pd 3d5/2 core level peak that corresponds to pure Pd NPs (335.5 eV) is shifted compared to bulk Pd (335.2 eV) which indicates the presence of NPs53 (see also Figure 4f). In the Ag32Au42Pd26 and Ag27Au17Pd56 samples, the Ag 3d5/2 core level peaks (at 368.3 and 368.0 eV, respectively) presented chemical shifts compatible with the formation of an alloy. The corresponding Pd 3d5/2 core level peaks (at 335.4 eV) do not change significantly with the chemical composition of the nanoparticles. In AgPd NP systems, a decrease in the BE of Pd 3d5/2 was attributed to electron donating effect from Ag to Pd.54 The positions of the Au 4f7/2 core level peaks are 84.3 and 84.0 eV for the Ag32Au42Pd26 and Ag27Au17Pd56 samples, respectively. These binding energies are close to the case of pure Au NPs (84.3 eV) and bulk Au (84 eV). These results evidenced the complexity of the analysis of trimetallic nanoparticled systems. Changes observed in the BE indicate the presence of alloyed NPs although a detailed analysis of the chemical shifts goes beyond the scope of this work. E. HRTEM Characterization of the Nanoparticles. Since the XPS technique provides an averaged chemical information,

Figure 8. (a) Ag 3d and Pd 3d core level spectra as a function of chemical composition in AuAgPd nanoparticles. (b) Au 4f core level spectra as a function of chemical composition in AuAgPd nanoparticles.

we have performed additional local chemical characterization using HRTEM and EDX techniques. Because of the difficulty to distinguish the signal coming from Ag and Pd in EDX, we present only the results that correspond to AgAu alloyed NPs. Figure 9a displays a representative low-magnification TEM image of an assembly of alloyed Ag55Au45 NPs deposited directly on a carbon coated TEM grid. Figure 9b displays a HRTEM image of a representative individual NP. Since crystallographic distances in Ag, Au, and corresponding alloys are very similar, the identification of an alloy through the structural analysis is not straightforward. We have therefore preferred to perform EDX analysis of the NPs. This is illustrated in Figure 9c where we show the evolution of the intensity of Au L and Ag L lines along the line profile indicated in Figure 9b. As can be clearly observed, the Au and Ag contents are homogeneous along the scanning line and hence in the particle. Because of the low EDX signal when performing the scanning (as a consequence of small NPs size), atomic concentration cannot be extracted from the scan. However, a point analysis (that allows a better statistics and hence a chemical determination) at the center of the nanoparticle 11246

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

In the case of Ag depletion in the whole NP surface, a core− shell like behavior would have been observed. In such a case, Au concentration would have decreased in the center of the NP,55 which is not the case for our nanoparticles. It is interesting to note that no Janus NPs could be observed, and the analysis of the HRTEM images revealed that the majority of the NPs were alloyed. Additionally, a reduced proportion of NPs presented a core−shell structure and a few of them were made of pure Au and Ag. Interestingly, the reduced number of core−shell and pure NPs presented smaller diameters that will allow the use of a quadrupole mass filter to select a given kind of NPs. Such refinement of the technique as well as a detailed study of the impact of the working parameters of the MICS on the relative proportion of each kind of NPs will be performed and presented in a forthcoming work that requires an in-depth HRTEM study of the fabricated NPs. Here we prefer to focus on the demonstration of the new capabilities of the MICS and show that alloyed NPs of controlled stoichiometry can be fabricated under UHV conditions.

IV. CONCLUSIONS We have presented a bottom-up route for the generation of nanoparticles. The method is based on a modified design of the ion cluster source where the single 2 in. magnetron has been replaced by three small magnetrons. It is shown that the individual control of the working parameters of the magnetrons allows the fabrication of nanoparticles with adjustable chemical composition. In particular, we have grown and characterized a series of high-purity bimetallic (Au−Ag) and trimetallic (Au− Ag−Pd) nanoparticles with controlled stoichiometry. The XPS data clearly show that the nanoparticles are made of AuAg or AgAuPd alloys with tunable stoichiometry while the HRTEM results show that the elements are homogeneously distributed over the nanoparticles.



AUTHOR INFORMATION

Corresponding Author

*E-mail [email protected]; tel +34 91 334 90 98; fax +34 91 372 06 23. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors acknowledge the Spanish Ministerio de Ciencia e Innovación and Comisión Interministerial para la Ciencia Y la Tecnologiá - CICYT under Contracts MAT2008-06765-C0202, MAT2011-29194-C02-02, MAT2009-08650, and CSD2007 00041 (Nanoselect) and through the FPI program for financial ́ support. The Consejo Superior de Investigaciones Cientificas is also acknowledged. D. Llamosa P. acknowledges financial support from Ministerio de Ciencia e Innovación under Contract JAEPre-09-01925. The HRTEM microscopy works have been conducted in the “Laboratorio de Microscopias Avanzadas” at “Instituto de Nanociencia de Aragon Universidad de Zaragoza”. Authors acknowledge the LMAINA for offering access to their instruments and expertise.

Figure 9. (a) Low-magnification TEM of an assembly of Ag55Au45 alloyed NPs. (b) HRTEM image of a representative Ag55Au45 alloyed NP. (c) Corresponding EDX profile performed at the K and L edges of Ag and Au along the line depicted in (b).

revealed that the atomic concentration in Ag51Au49, in good agreement with the chemical composition determined by XPS. Note that the relatively low electron beam current (the reduction of the electron beam current is mandatory otherwise the NPs are destroyed by the electron beam) used when performing the profile scans induces some scattering in the profile lines like those presented in Figure 9c. The key point in the profiles displayed in Figure 9c is that the concentrations of Au and Ag present the highest values in the middle of the NP.



REFERENCES

(1) Schmelzer, J., Jr.; Brown, S. A.; Wurl, A.; Hyslop, M.; Blaikie, R. J. Finite-size effects in the conductivity of cluster assembled nanostructures. Phys. Rev. Lett. 2002, 88, 226802−22685. (2) Russo, R.; Cianchi, A.; Akhmadeev, Y. H.; Catani, L.; Langner, J.; Lorkiewicz, J.; Polini, R.; Ruggiero, B.; Sadowski, M. J.; Tazzari, S.;

11247

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

Koval, N. N. UHV arc for high quality film deposition. Surf. Coat. Technol. 2006, 201, 3987−3992. (3) Russo, R.; Catani, L.; Cianchi, A.; Tazzari, S.; Langner, J. Supercond. Sci. Technol. 2005, 18, L41−L44. (4) Granqvist, C. G.; Buhrman, R. A. Ultrafine metal particles. J. Appl. Phys. 1976, 47, 2200−2219. (5) Sattler, K.; Mühlbach, J.; Recknagel, E. Generation of metal clusters containing from 2 to 500 atoms. Phys. Rev. Lett. 1980, 45, 821−824. (6) Mühlbach, J.; Pfau, P.; Recknagel, E.; Sattler, K. Cluster emission from the surfaces of Bi, Sb and Se. Surf. Sci. 1981, 106, 18−26. (7) Milani, P.; deHeer, W. A. Improved pulsed laser vaporization source for production of intense beams of neutral and ionized clusters. Rev. Sci. Instrum. 1990, 61, 1835−1838. (8) Siekmann, H. R.; Lüder, Ch.; Faehrmann, J.; Lutz, H. O.; Meiwes-Broer, K. H. The pulsed arc cluster ion source (PACIS). Z. Phys. D: At., Mol. Clusters 1991, 20, 417−420. (9) Haberland, H.; Karrais, M.; Mall, M. A new type of cluster and cluster ion source. Z. Phys. D: At., Mol. Clusters 1991, 20, 413−415. (10) Haberland, H.; Karrais, M.; Mall, M.; Thurner, Y. Thin films from energetic cluster impact: A feasibility study. J. Vac. Sci. Technol., A 1992, 10, 3266−3271. (11) Haberland, H.; Mall, M.; Moseler, M.; Qiang, Y.; Reiners, T.; Thurner, Y. Filling of micron sized contact holes with copper by energetic cluster impact. J. Vac. Sci. Technol., A 1994, 12, 2925−2930. (12) Xirouchaki, C.; Palmer, R. E. Deposition of size-selected metal clusters generated by magnetron sputtering and gas condensation: a progress review. Philos. Trans. R. Soc., A 2004, 362, 117−124. (13) Bansmann, J.; Baker, S. H.; Binns, C.; Blackman, J. A.; Bucher, J.-P.; Dorantes-Dávila, J.; Dupuis, V.; Favre, L.; Kechrakos, D.; Kleibert, A.; Meiwes-Broer, K.-H.; Pastor, G. M.; Perez, A.; Toulemonde, O.; Trohidou, K. N.; Tuaillon, J.; Xie, Y. Magnetic and structural properties of isolated and assembled clusters. Surf. Sci. Rep. 2005, 56, 189−275. (14) Wegner, K.; Piseri, P.; Vahedi Tafreshi, H.; Milani, P. Cluster beam deposition: a tool for nanoscale science and technology. J. Phys. D: Appl. Phys. 2006, 39, R439−R459. (15) Binns, C.; Trohidou, K. N.; Bansmann, J.; Baker, S. H.; Blackman, J. A.; Bucher, J.-P.; Kechrakos, D.; Kleibert, A.; Louch, S.; Meiwes-Broer, K.-H.; Pastor, G. M.; Perez, A.; Xie, Y. The behaviour of nanostructured magnetic materials produced by depositing gasphase nanoparticles. J. Phys. D: Appl. Phys. 2005, 38, R357−R379. (16) Perez, A.; Mélinon, P.; Dupuis, V.; Bardotti, L.; Masenelli, B.; Tournus, F.; Prével, B.; Tuaillon-Combes, J.; Bernstein, E.; Tamion, A.; Blanc, N.; Taïnoff, D.; Boisron, O.; Guiraud, G.; Broyer, M.; Pellarin, M.; Del Fatti, N.; Vallée, F.; Cottancin, E.; Lermé, J.; Vialle, J. L.; Bonnet, C.; Maioli, P.; Crut, A.; Clavier, C.; Rousset, J. L.; Morfin, F. Functional nanostructures from clusters. Int. J. Nanotechnol. 2010, 7, 523−574. (17) Oxford Applied Research, http://www.oaresearch.co.uk. (18) Baker, S. H.; Thornton, S. C.; Keen, A. M.; Preston, T. I.; Norris, C.; Edmonds, K. W.; Binns, C. The construction of a gas aggregation source for the preparation of mass-selected ultrasmall metal particles. Rev. Sci. Instrum. 1997, 68, 1853−1857. (19) Baker, S. H.; Thornton, S. C.; Edmonds, K. W.; Maher, M. J.; Norris, C.; Binns, C. The construction of a gas aggregation source for the preparation of size-selected nanoscale transition metal clusters. Rev. Sci. Instrum. 2000, 71, 3178−3183. (20) Imanaka, M.; Katayama, T.; Ohshiro, Y.; Watanabe, S.; Arai, H.; Nakagawa, T. Nanocluster ion source by plasma-gas aggregation. Rev. Sci. Instrum. 2004, 75, 1907−1909. (21) Iles, G. N.; Baker, S. H.; Thornton, S. C.; Binns, C. Enhanced capability in a gas aggregation source for magnetic nanoparticles. J. Appl. Phys. 2009, 105, 024306−7. (22) Momin, T.; Bhowmick, A. A new magnetron based gas aggregation source of metal nanoclusters coupled to a double time-offlight mass spectrometer system. Rev. Sci. Instrum. 2010, 81, 075110− 10.

(23) Stoyanov, S.; Huang, Y.; Zhang, Y.; Skumryev, V.; Hadjipanayis, G. C.; Weller, D. Fabrication of ordered FePt nanoparticles with a cluster gun. J. Appl. Phys. 2003, 93, 7190−3. (24) Qiu, J.-M.; Xu, Y.-H.; Judy Jack, H.; Wang, J.-P. Nanocluster deposition for high density magnetic recording tape media. J. Appl. Phys. 2005, 97, 10P704−3. (25) Bai, J.; Wang, J.-P. High-magnetic-moment core-shell-type FeCo−Au/Ag nanoparticles. Appl. Phys. Lett. 2005, 87, 152502−3. (26) Kennedy, M. K.; Kruis, F. E.; Fissan, H.; Mehta, B. R.; Stappert, S.; Dumpich, G. Tailored nanoparticle films from monosized tin oxide nanocrystals: Particle synthesis, film formation, and size-dependent gas-sensing properties. J. Appl. Phys. 2003, 93, 551−560. (27) Fonzo, F. Di; Gidwani, A.; Fan, M. H.; Neumann, D.; Iordanoglou, D. I.; Heberlein, J. V. R.; McMurry, P. H.; Girshick, S. L.; Tymiak, N.; Gerberich, W. W.; Rao, N. P. Focused nanoparticle-beam deposition of patterned microstructures. Appl. Phys. Lett. 2000, 77, 910−3. (28) Goldby, I. M.; Issendorff, B. von; Kuipers, L.; Palmer, R. E. Gas condensation source for production and deposition of size-selected metal clusters. Rev. Sci. Instrum. 1997, 68, 3327−8. (29) Majumdar, A.; Köpp, D.; Ganeva, M.; Datta, D.; Bhattacharyya, S.; Hippler, R. Development of metal nanocluster ion source based on dc magnetron plasma sputtering at room temperature. Rev. Sci. Instrum. 2009, 80, 095103−6. (30) Pellarin, M.; Ray, C.; Lermé, J.; Vialle, J. L.; Broyer, M.; Mélinon, P. Gas phase study of silicon−C60 complexes: Surface coating and polymerization. J. Chem. Phys. 2000, 112, 8436−8445. (31) Masenelli, B.; Tournus, F.; Mélinon, P.; Blasé, X.; Perez, A.; Pellarin, M.; Broyer, M. Nanostructured films from (C60)(n)Si-m clusters. Appl. Surf. Sci. 2004, 226, 226−230. (32) Sumiyama, K.; Hihara, T.; Liang Peng, D.; Katoh, R. Structure and magnetic properties of Co/CoO and Co/Si core−shell cluster assemblies prepared via gas-phase. Sci. Technol. Adv. Mater. 2005, 6, 18−26. (33) Quesnel, E.; Pauliac-Vaujour, E.; Muffato, V. Modeling metallic nanoparticle synthesis in a magnetron-based nanocluster source by gas condensation of a sputtered vapor. J. Appl. Phys. 2010, 107, 054309−8. (34) Toshima, N. Capped Bimetallic and Trimetallic Nanoparticles for Catalysis and Information Technology. Macromol. Symp. 2008, 279, 27−39. (35) Zhong, C.-J.; Luo, J.; Fang, B.; Wanjala, B. N.; Njoki, P. N.; Loukrakpam, R.; Yin, J. Nanostructured catalysts in fuel cells. Nanotechnology 2010, 21, 062001−20. (36) Román García, E. L.; Martínez Orellana, L.; Díaz Lagos, M.; Huttel, Y. Dispositivo y procedimiento de fabricación de nanó particulas. Spanish Patent P201030059, PCT/ES2011/070032, 2010. (37) Nanotec Electrónica S.L., http://www.nanotec.es. (38) Martínez, L.; Tello, M.; Díaz, M.; Román, E.; Garcia, R.; Huttel, Y. Aspect-ratio and lateral-resolution enhancement in force microscopy by attaching nanoclusters generated by an ion cluster source at the end of a silicon tip. Rev. Sci. Instrum. 2011, 82, 023710−7. (39) Román García, E. L.; Martínez Orellana, L.; Díaz Lagos, M.; Huttel, Y. Modificación de puntas de microscopiá de fuerzas atómicas ́ con una fuente de agregados. mediante depósito de nanoparticulas Spanish Patent P201030712, PCT/ES2011/070319, 2010. (40) Horcas, I.; Fernández, R.; Gómez-Rodríguez, J. M.; Colchero, J.; Gómez-Herrero, J.; Baró, A. M. WSxM: A software for scanning probe microscopy and a tool for nanotechnology. Rev. Sci. Instrum. 2007, 78, 013705−8. (41) Colino García, J. M. Unidad de pulverización catódica de blancos circulares. Spanish Patent P200900929, 2009. (42) Roldan Cuenya, B.; B. Synthesis and catalytic properties of metal nanoparticles: Size, shape, support, composition, and oxidation state effects. Thin Solid Films 2010, 518, 3127−3150. (43) Liu, C.; Zou, B.; Rondinone, A. J.; Zhang, Z. J. Chemical Control of Superparamagnetic Properties of Magnesium and Cobalt Spinel Ferrite Nanoparticles through Atomic Level Magnetic Couplings. J. Am. Chem. Soc. 2000, 122, 6263−6267. 11248

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249

Langmuir

Article

(44) Fernández-Martinez, I.; Martín-González, M. S.; GonzálezArrabal, R.; Á lvarez-Sánchez, R.; Briones, F.; Costa-Krämer, J. L. Nitrided FeB amorphous thin films for magneto mechanical systems. J. Magn. Magn. Mater. 2008, 320, 68−75. (45) O’Grady, K.; Bradbury, A. Particle size analysis in ferrofluids. J. Magn. Magn. Mater. 1983, 39, 91−94. (46) Díaz, M.; Martínez, L.; Ruano, M.; Llamosa, P. D.; Román, E.; García-Hernández, M.; Ballesteros, C.; Fermento, R.; Cebollada, A.; Armelles, G.; Huttel, Y. Morphological, structural, and magnetic properties of Co nanoparticles in a silicon oxide matrix. J. Nanopart. Res. 2011, 13, 5321−5333. (47) Harikumar, K. R.; Ghosh, S.; Rao, C. N. R. X-ray Photoelectron Spectroscopic Investigations of Cu-Ni, Au-Ag, Ni-Pd, and Cu-Pd Bimetallic Clusters. J. Phys. Chem. A 1997, 101, 536−540. (48) Wang, A.-Q.; Liu, J.-H.; Lin, S. D.; Lin, T.-S.; Mou, C.-Y. A novel efficient Au−Ag alloy catalyst system: preparation, activity, and characterization. J. Catal. 2005, 233, 186−197. (49) Lim, D. C.; Lopez-Salido, I.; Dietsche, R.; Dok Kim, Y. Interactions of oxygen and CO with AgAu bimetallic nanoparticles on sputtered highly ordered pyrolytic graphite (HOPG) surfaces. Surf. Sci. 2007, 601, 5635−5642. (50) Lopez-Salido, I.; Lim, D. C.; Dietsche, R.; Bertram, N.; Dok Kim, Y. Electronic and Geometric Properties of Au Nanoparticles on Highly Ordered Pyrolytic Graphite (HOPG) Studied Using X-ray Photoelectron Spectroscopy (XPS) and Scannong Tunneling Microscopy (STM). J. Phys. Chem. B 2006, 110, 1128−1136. (51) Sellinger, A. T.; Aburada, T.; Fitz-Gerald, J. M. Synthesis of multimetallic nanoparticles using a solution-based pulsed laser deposition approach. Proc. SPIE 2008, 7005, 700516. (52) Venkatesan, P.; Santhanalakshmi, J. Designed Synthesis of Au/ Ag/Pd Trimetallic Nanoparticle-Based Catalysts for Sonogashira Coupling Reactions. Langmuir 2010, 26, 12225−12229. (53) Aruna, I.; Mehta, B. R.; Malhotra, L. K.; Shivaprasad. Size dependence of core and valence binding energies in Pd nanoparticles: Interplay of quantum confinement and coordination reduction. S. M. J. Appl. Phys. 2008, 104, 064308−064312. (54) He, J.; Ichinose, I.; Kunitake, T.; Nakao, A.; Shiraishi, Y.; Toshima, N. Facile Fabrication of Ag-Pd Bimetallic Nanoparticles in Ultrathin TiO2-Gel Films: Nanoparticle Morphology and Catalytic Activity. J. Am. Chem. Soc. 2003, 125, 11034−11040. (55) Thomas, J.; Gemming, T. Shells on nanowires detected by analytical TEM. Appl. Surf. Sci. 2005, 252, 245−251.

11249

dx.doi.org/10.1021/la3022134 | Langmuir 2012, 28, 11241−11249